Complex Formation by Electrostatic Interaction between Carboxyl

Dendrimers Derived from 1 → 3 Branching Motifs ... Langmuir 0 (proofing), .... EPL (Europhysics Letters) 2010 92, 18001 ... Polyelectrolyte complexe...
0 downloads 0 Views 152KB Size
Langmuir 1999, 15, 4245-4250

4245

Complex Formation by Electrostatic Interaction between Carboxyl-Terminated Dendrimers and Oppositely Charged Polyelectrolytes† Nobuhiro Miura,‡ Paul L. Dubin,* C. N. Moorefield,§ and G. R. Newkome§ Department of Chemistry, Indiana UniversitysPurdue University, Indianapolis, Indianapolis, Indiana 46202 Received February 8, 1999. In Final Form: March 22, 1999 Complex formation between a carboxyl-terminated cascade polymer (generation 3) and several cationic polyelectrolytes of varying linear charge density was studied as a function of ionic strength, by turbidimetric titration and dynamic light scattering. Tetramethylammonium chloride was used to adjust the ionic strength in order to avoid sodium counterion binding to dendrimer carboxyl groups. Complex formation occurred abruptly at a critical pH, as signaled by a sudden change in either the turbidity or the apparent Stokes radius from dynamic light scattering. The pHc of incipient complex formation was converted to the critical surface charge density σc. Under conditions of low or moderate ionic strength (I), it was confirmed that σc is roughly proportional to κ/ξ, where κ ∼ I1/2 is the Debye-Hu¨ckel parameter and ξ is the linear charge density of the polyelectrolyte.

Introduction Intermacroionic complex formation is a basic phenomenon in biological systems1 as well as in a number of technological processes;2 however, in the case of biological systems such as protein-DNA and enzyme-substrate complexes, the mechanism of intermolecular interaction is difficult to fully elucidate in physicochemical terms. For example, ligand binding constants, related to biological activity, cannot be quantitatively explained solely on the basis of fundamental variables such as macromolecular charge density or the electrostatic screening effect of the counterion. For such reasons, and also because of the technological relevance of colloid-polymer interactions, we focus on complex formation arising from electrostatic interactions in synthetic systems of polyelectrolytes and oppositely charged colloid particles. The colloidal particle in this study is a symmetric, uniform, and spherical dendrimer3 which provides an ideal model for small charged colloids. The binding of polyelectrolytes and oppositely charged particles depends predominantly on three variables: the ionic strength (I), the polymer linear charge density (ξ), and the colloid surface charge density (σ). Phase-transition-like behavior for the interaction of polyelectrolytes with oppositely charged particles has been predicted by several theoretical investigations.4-8 Previous experi†

Presented at Polyelectrolytes ’98, Inuyama, Japan, May 31June 3, 1998. ‡ Current address: Polymer Science & Engineering, University of Massachusetts, Amherst, MA 01002. § Department of Chemistry, University of South Florida, Tampa, FL 33620. (1) Lilley, D. M. J. DNA-Protein: Structural Interactions; IRL Press: Oxford, U.K., 1995; pp 26-27. (2) Evans, D. F.; Wennerstrom, H. The Colloidal Domain: Where Physics, Chemistry, Biology, and Technology Meet; VCH Publishers: New York, 1994; pp 455-496. (3) Newkome, G. R.; Moorefield, C. N.; Vo¨gtle, F. Dendritic Molecules: Concepts, Syntheses, Perspectives; VCH: Weinheim, Germany, 1996. (4) Odijk, T. Langmuir 1991, 7, 1991. (5) von Goeler, F.; Muthukumar, M. J. Chem. Phys. 1994, 100, 7796. (6) Wiegel, F. W. J. Phys. A: Math. Gen. 1977, 10, 299. (7) Varoqui, R.; Johner A.; Elaissari, A. J. Chem. Phys. 1991, 94, 6874.

mental studies on micelle-,9-12 dendrimer-,13,14 and protein-polyelectrolyte15-17 systems have revealed the existence of such critical conditions for complexation. Studies with polyelectrolytes and oppositely charged micelles have suggested the empirical relationship

σcξ ∼ κ

(1)

where κ is the Debye-Hu¨ckel parameter (κ ∼ I1/2) and σc is the critical surface charge density of the micelle at the point of incipient complex formation.9 In the case of a dendrimer-polyelectrolyte system, it was observed that σc depends on the ionic strength,13,14 but the relationship among σc, ξ, and κ was not confirmed. In this report, we examine the dependence of σc on ξ and κ. The quantity ξ is varied by using copolymers of ((methacrylamido)propyl)trimethylammonium chloride (MAPTAC) and acrylamide (AAm). Sodium counterion binding to carboxyl-terminated dendrimer has been demonstrated by potentiometric titration,18 so it may be expected to affect complexation; therefore, both TMACl and NaCl were used to adjust the ionic strength (I). The mechanism of phase separation is also briefly discussed in this study, since observations of phase separation by complexation may help us to understand the mechanism (8) Evers, O. A.; Fleer, G. J.; Scheutjens, J. M. H. M.; Lyklema, J. J. Colloid Interface Sci. 1986, 111, 446. (9) McQuigg, D. W.; Kaplan, J. I.; Dubin, L. P. J. Phys. Chem. 1992, 96, 1973. (10) Li, Y.; Xia, J.; Dubin, P. L. Macromolecules 1994, 27, 7049. (11) Li, Y.; Dubin, L. P.; Dautzenberg, H.; Luck, U.; Hartmann, J.; Tuzar, Z. Macromolecules 1995, 28, 6795. (12) Yoshida, K.; Sokhakian, S.; Dubin, P. L. Colloids Surf., in press. (13) Li, Y.; Dubin, P. L.; Spindler, R.; Tomalia, D. A. Macromolecules 1995, 28, 8426. (14) Zhang, H.; Dubin, P. L.; Spindler, R.; Tomalia, D. A. Ber. BunsenGes. Phys. Chem. 1996, 100, 923. (15) Li, Y.; Mattison, K. W.; Dubin, P. L.; Havel, H. A.; Edwards, S. L. Biopolymers 1996, 38, 527. (16) Xia, J.; Mattison, K.; Romano, V.; Muhoberac, B. B. Biopolymers 1997, 41, 359. (17) Mattison, K. W.; Dubin, P. L.; Brittain, I. J. J. Phys. Chem. B 1998, 102, 3830. (18) Zhang, H.; Dubin, P. L.; Kaplan, J.; Moorefield, C. N.; Newkome, G. R. J. Phys. Chem. 1997, 101, 3495.

10.1021/la990125l CCC: $18.00 © 1999 American Chemical Society Published on Web 05/13/1999

4246 Langmuir, Vol. 15, No. 12, 1999

Miura et al.

of precipitation and complex coacervation (liquid-liquid polyelectrolyte phase separation) in general. Experimental Section Materials. The third-generation cascade methane[4]-(3-oxo6-oxa-2-azaheptylidyne)-(3-oxo-2-azapentylidyne)2-propanoic acid (G3) has 108 terminal COOH groups and a theoretical molecular weight of 12 345.19,20 Poly(diallyldimethylammonium chloride) (PDADMAC), molecular weight 458 000, was synthesized by free-radical polymerization of diallyldimethylammonium chloride, purified, and characterized in the laboratory of Dr. W. Jaeger (FraunhoferInstitut, Teltow, Germany). Poly[(methacrylamido)propyl)trimethylammonium chloride] (PMAPTAC) was a gift from Clairol corporation, Stamford, CT. Copolymers of MAPTAC and acrylamide (AAm), 60, 40, and 30% MAPTAC, were prepared by Dr. Takeshi Sato at Osaka University. A 15% MAPTAC and AAm copolymer was from Jefferson Chemical Company, Inc., Bellaire, Texas. Studies with micelle-polyelectrolyte or protein-polyelectrolyte systems have shown that polymer MW has no effect on σc,10,11 so no attempt was made to characterize MW for these copolymers. The average linear polymer charge density ξ was obtained as the reciprocal of contour length charge spacing: 2.5 Å for PMAPTAC, 6.2 Å for PDADMAC, 3.1 Å for MAPTACAAm (80:20), 4.2 Å for MAPTAC-AAm (60:40), 6.3 Å for MAPTAC-AAm (40:60), and 16.7 Å for MAPTAC-AAm (15:85). Sodium chloride (NaCl) and tetramethylammonium chloride (TMACl) were purchased from Sigma and Aldrich Chemical Co. Inc., respectively. NaOH, HCl, and buffer solutions of pH 4.00 ( 0.01 and pH 7.00 ( 0.01 were from Fisher Scientific. Potentiometric Titration. Potentiometric titrations with 0.5 M HCl or 0.5 M NaOH were performed with an Orion pH meter 811 at 23 ( 1 °C. The pH meter was calibrated with 4.00 ( 0.01 and 7.00 ( 0.01 buffers just before titration. The instrumental drift was less than 0.02 during titration. Nitrogen gas was bubbled through the solution to remove carbon dioxide in the solution for 10 min before the titration, and the solution was then maintained under N2. A 2.0 mL microburet was used to add 0.500 M NaOH or 0.500 M HCl to a 10.00 mL aqueous solution of dendrimer (0.50 g/L), containing the desired concentration of NaCl or TMACl and with initial pH adjusted to 7.00 ( 0.03. These titrations were always accompanied by a blank titration of dendrimer-free solution. The amount of acid or base contributed to the neutralization reaction was estimated as the difference, ∆V, in added HCl or NaOH between dendrimer solution and blank solution at the same pH. Successful blank correction was indicated by the appearance of asymptotic end points corresponding to degrees of dissociation R ) 0 and 1 in the ∆V versus pH plots (see Figure 1). R was determined as

R ) (∆V - ∆V0)/∆VT ) [COO-]/([COOH] + [COO-])

(2)

where ∆VT is the total value between R ) 0 and 1, and ∆V0 is ∆V at R ) 0. Sample plots of pH against the R value of the dendrimer in 0.1 M TMACl and NaCl are shown in Figure 2. Turbidimetric Titration. The pH dependence of turbidity was measured using a Brinkman PC 800 colorimeter equipped with a 1 cm path length optical probe, at 23 ( 1 °C. The colorimeter was calibrated to read 100% transmittance with Milli-Q water. The solutions (0.10 g/L) were filtered with 0.2 µm Whatman filters before turbidimetric titration. All titrations were carried out with magnetic stirring, and the time interval between measurements was about 1 min. After mixing, solutions with transmittances corresponding to 100 - %T < 5 were always stable with respect to both turbidity and DLS measurements. Dynamic Light Scattering. Dynamic light scattering (DLS) measurements were performed using a DynaPro-801 (Protein Solutions Inc., Charlottesville, VA), equipped with a 780 nm solidstate laser of approximately 25 mW power and an avalanche photodiode detector. All solutions were filtered with 0.2 µm (19) Newkome, G. R.; Young, J. K.; Baker, G. R.; Potter, R. L.; Audoly, L.; Cooper, D.; Weis, C. D. Macromolecules 1993, 26, 2394. (20) Young, J. K.; Baker, G. R.; Newkome, G. R.; Morris, K. F.; Johnson, C. S., Jr. Macromolecules 1994, 27, 3464.

Figure 1. Titration of dendrimer G3 in 0.1 M TMACl presented as blank-corrected titrant volume versus pH. Negative values of ∆V correspond to titration with 0.500 M NaOH, and positive values to titration with 0.500 M HCl. ∆VT is the total value between R ) 0 and R ) 1, and ∆V0 is ∆V at R ) 0.

Figure 2. G3 titration curves in TMACl: 0.050 M (9); 0.100 M (0); 0.200 M (b) 0.500 M (O); obtained from data as in Figure 1. Inset: G3 titration curves in 0.1 M TMACl (2) and NaCl (O). Whatman filters before DLS measurement. The intensityintensity autocorrelation function G(2)(τ), which expresses the temporal fluctuation of the intensity of scattered light, is obtained from measured light scattering by

G(2)(τ) )

∑I(t ) I(t - τ) j

j

(3)

where I(t) is the scattering intensity at time t and τ is the delay time. The autocorrelation function was analyzed via CONTIN. The decay constant τ0 was determined by fitting of G(2)(τ) to the function

G(2)(τ) ) B + C exp(τ/τ0)

(4)

where B is the baseline and C is the gain of the decay function. Assuming that the particle is uniform and that interparticle interactions are negligible, the translational diffusion coefficient of the molecule DT was estimated by

DT ) q2/τ0

(5)

in which q ) (4πn/λ0) sin(θ/2) is the magnitude of the scattering vector, n is the refractive index of the solution, θ is the scattering angle, and λ0 is the wavelength of light in a vacuum. Then the z-average apparent hydrodynamic radius RSapp of the particles was derived from DT using the Stokes-Einstein equation:

DT ) kT/6πηRSapp

(6)

Dendrimers and Oppositely Charged Polyelectrolytes

Figure 3. pKa versus R for G3 in TMACl: 0.050 M (9); 0.100 M (0); 0.200 M (b) 0.500 M (O). where k is the Boltzmann constant, T is the absolute temperature, and η is the solvent viscosity. Our interest is only the determination of the pH at which RSapp changes abruptly, so its absolute value is not of primary concern.

Results and Discussion Potentiometric Titration. The relationship between pH and R for dendrimer G3 in 0.1 M TMACl and NaCl is shown in Figure 2. The results in TMACl and NaCl converge below pH 5, but the slope in NaCl is steeper than that in TMACl in the region between pH 5 and pH 7. This result suggests that the binding of Na+ to dendrimer increases the acidity of the carboxylic acid groups, an effect not seen for the nonbinding TMA+. The tendency of Na+ to bind to carboxylate has been welldocumented.21 To avoid counterion binding, TMACl was therefore used to control ionic strength in addition to NaCl. Dendrimer solutions of 0.5 g/L at various ionic strengths of TMACl were titrated with 0.500 ( 0.002 M HCl and NaOH. Titration curves of G3 in various TMACl concentrations are shown in Figure 3 as the R-dependence of the apparent pK, pKa ≡ pH + log(1 - R)/R. Phase Separation and Complexation between Dendrimer G3 and Polyelectrolyte. Turbidimetric titrations for dendrimer G3 in the presence of MAPTACAAm (40%/60%) copolymer in TMACl at various ionic strengths are shown in Figure 4A. The sudden increase in turbidity at a well-defined pHc arises because phase separation is a result of complexation. The increase in pHc with ionic strength is due to screening. The behavior of PDADMAC-G3 in NaCl in Figure 4B is more complicated and will be discussed below. All the results are summarized in Tables 1 and 2 and plotted as pHc versus I in Figure 5. Since it is difficult to estimate Rc in the low-pH region, most of the data in Figure 5 are for the low-charge-density polyelectrolyte, and there is only one data point for PMAPTAC. The pH dependence of the apparent hydrodynamic radius RSapp for G3/PDADMAC in 0.50 M NaCl and for G3/MAPTAC-AAm (15%/85%) in 0.10 M NaCl is shown in Figure 6. For the copolymer, the discontinuity in RSapp corresponds to the pHc determined by turbidimetric titration and confirms the use of the latter method for determining pHc. However, in the case of PDADMAC, RSapp (21) (a) Gregor, H. P. In Polyelectrolytes; Selegny, E., Ed.; D. Reidel: Dordrecht, 1974; p 87. (b) Struass, U. P.; Leung, Y. P. J. Am. Chem. Soc. 1965, 87, 1476. (c) Rinaudo, M.; Milas, M. J. Chim. Phys. 1969, 76, 254. (d) Eldridge, R. J.; Treloar, F. E. J. Phys. Chem. 1976, 80, 1513.

Langmuir, Vol. 15, No. 12, 1999 4247

Figure 4. (A) Turbidity (100 - %T) versus pH for dendrimer G3 + PMAPTAC-AAm (40%/60%) at the ionic strengths (TMACl) 0.100 M (0), 0.200 M (b), 0.300 M (9), and 0.400 M (O). (B) Turbidity (100 - %T) versus pH for dendrimer G3 + PDADMAC at the ionic strengths (NaCl) 0.200 M (9), 0.350 M (b) and 0.500 M (O).

changes gradually, and the discontinuity is not clear and does not agree with the pHc determined by turbidimetric titration. The effects of NaCl versus TMACl may also be seen for dendrimer G3/MAPTAC-AAm (60/40) from the plots of pHc versus I in Figure 5 (open and filled triangles). The critical pH in TMACl is smaller than that in NaCl, this tendency increasing with increasing ionic strength. These results suggest that Na+ binding to dendrimer interferes with its binding to polyelectrolyte and thus impedes complexation and phase separation. As seen in Figure 4B for PDADMAC in NaCl, pHc is separated from pHφ, corresponding to the formation of a soluble complex without phase separation. The maximum in turbidity for PDADMAC in 0.50 M NaCl as a function of pH may in fact correspond to the formation of soluble complexes, followed by their redissolution at high pH due to charge reversal. Evidence of soluble complex formation may be discerned also for the other homopolymer, PMAPTAC, as seen in Figure 7. The “steps” between pHc and pHφ in Figure 7 result from the instrumental limit of (0.1%T; however, the instrumental drift of the system is less than 0.1 per 1 h, so the subtle increase in turbidity from pH 3.5 to pH 4.1 is real. The pH dependence of RSapp for PDADMAC in Figure 6 indicates the presence of soluble complexes before phase separation. Such soluble complexes were also observed in protein-polyelectrolyte and micelle-polyelectrolyte systems. Since phase separation is likely to occur by charge neutralization,22 soluble complexes presumably bear some excess charge. The situation is represented schematically in Figure 8. When complex formation occurs at low R, the bound dendrimer charge cannot cancel the highly charged polyelectrolyte and excess charge arises. Thus, soluble complex is observed in systems showing low pHc. Although the linear charge density ξ of PDADMAC is nearly the same as that of the 40% MAPTAC copolymer, its pHc is lower. This shift of pHc is too large to explain by the difference between Na+ and TMA+, since the effect of counterion binding is small at low ionic strength. Similar anomalous behavior has been observed in proteinpolyelectrolyte systems (e.g. bovine serum albumin with PDADMAC versus a 50/50 copolymer of trimethylaminoethylacrylate (CMA) with acrylamide)23 and in micelle(22) Kruyt, H. R. Colloid Science; Elsevier Publication Company: New York, 1949; Vol. II, pp 335-432.

4248 Langmuir, Vol. 15, No. 12, 1999

Miura et al.

Table 1. pHc, rc, σg, σGC, and ξσGC for PMAPTAC and Copolymers of MAPTAC and Acrylamide Obtained from Turbidimetric Titration at Various Ionic Strengths MAPTAC (%)

ξ (×108 cm-1)

I

pHc

Rc

σc, (×10-4 esu/cm2)

σc,GC (×10-4 esu/cm2)

σc,GCξ (×103 esu/cm3)

100 60

0.40 0.24

40

0.16

30

0.12

15

0.06

0.30 0.10 0.20 0.30 0.40 0.10 0.20 0.30 0.40 0.10 0.20 0.30 0.025 0.050 0.075 0.100

3.44 3.46 3.77 4.24 4.60 3.47 4.02 4.52 4.96 3.65 4.36 4.93 4.30 4.66 4.92 5.25

0.08 0.07 0.10 0.17 0.22 0.07 0.15 0.21 0.30 0.08 0.19 0.30 0.13 0.14 0.23 0.30

1.1 1.0 1.4 2.3 3.0 1.0 2.1 2.9 4.1 1.1 2.6 4.1 1.8 1.9 3.2 4.1

0.7 0.7 1.2 1.8 2.1 0.7 1.6 2.0 2.6 0.9 2.1 2.6 1.29 1.89 2.24 2.65

3.0 1.7 2.9 4.4 5.1 1.1 2.6 3.2 4.1 1.0 2.5 3.1 0.77 1.13 1.34 1.57

Figure 6. pH dependence of the apparent Stokes-Einstein radius for G3 + PDADMAC in 0.5 M NaCl (O) and G3PMAPTAC-AAm (15%/85%) in 0.1 M NaCl (b). Figure 5. pHc versus ionic strength in TMACl for PMAPTAC (O), MAPTAC-AAm copolymer (60%/40%) (4), MAPTACAAm copolymer (40%/60%) (0), MAPTAC-AAm copolymer (30%/70%) (3), MAPTAC-AAm copolymer (15%/85%) (]), MAPTAC-AAm copolymer (60%/40%) in NaCl (2), and PDADMAC in NaCl (b). Table 2. pHc, rc, σg, σGC, and ξσGC for PMAPTAC (ξ ) 0.16 × 108 cm-1) Obtained from Turbidimetric Titration at Various Ionic Strengths I

pHc

Rc

σg, (×10-4 esu/cm2)

σc,GC (×10-4 esu/cm2)

σc,GC (×10-3 esu/cm2)

0.20 0.30 0.35 0.40 0.45 0.50

3.10 3.42 3.38 3.58 3.72 3.7

0.03 0.05 0.06 0.08 0.09 0.1

0.4 0.7 0.8 1.1 1.2 1.4

0.2 0.3 0.4 0.5 0.5 0.5

0.3 0.5 0.6 0.8 0.8 0.8

polyelectrolyte systems (Triton X-100/sodium dodecyl sulfate with PDADMAC versus PMAPTAC).24 In both cases, binding of anionic colloids to polycations was stronger for PDADMAC. One explanation may be to consider the polyelectrolyte as a cylinder, whose surface charge density then depends on the side chain length. Since the distances between the backbone and the side chain charge are about 6 Å for PMAPTAC and about 2 Å for PDADMAC, the value 0.008 (electronic charges per (23) Park, J. M.; Muhoberac, B. B.; Dubin, P. L.; Xia, J. Macromolecules 1992, 25, 290. (24) Zhang, H.; et al. Unpublished.

Figure 7. Plots of turbidimetric titration for G3 + PMAPTAC at 0.3 M NaCl (O) and TMACl (b).

Å-2) for the cylinder geometric surface charge density obtained for PDADMAC is considerably larger than the value 0.005 obtained for the CMA-AAm copolymer, whose linear charge density is nearly the same as that of the 50% MAPTAC-AAm copolymer.23 Consequently, G3 binding is stronger for PDADMAC than for MAPTACAAm copolymers. Critical Surface Charge Density. The proton dissociation constant K is defined as

Kapp ) [COO-][H+]/[COOH] ) R[H+]/(1 - R) (7)

Dendrimers and Oppositely Charged Polyelectrolytes

Langmuir, Vol. 15, No. 12, 1999 4249

Figure 8. Schematic of complex formation for dendrimer with (a) high charged polyelectrolyte and (b) low charged polyelectrolyte.

If the electrostatic free energy required to dissociate one mole of protons at a given degree of dissociation is ∆Gel(R), then25

pKapp ) pK0 + 0.434∆Gel(R)/RT

(8)

where R is the gas constant and K0 is the characteristic dissociation constant when electrostatic interactions with other dissociated groups are absent. pK0 is generally obtained as pKapp in the limit of R ) 0. Since ∆Gel(R) is the work of transferring a proton from the bulk solution to the dendrimer surface, then if eNψ0 ) ∆Gel(R),

ψ0 ) 2.303(pKa - pK0)kT/e

Figure 9. σc vs κ in TMACl for: PMAPTAC (O), MAPTACAAm copolymer (60%/40%) (4), MAPTAC-AAm copolymer (40%/60%) (0), MAPTAC-AAm copolymer (30/70%) (3), MAPTAC-AAm copolymer (15%/85%) (]), MAPTAC-AAm copolymer (60%/40%) in NaCl (2), and DMDAO-PAMPS in NaCl (+).

(9)

where e is the elementary charge, 4.8 × 10-10 esu, N is Avogadro’s number, and ψ0 is the surface potential (statvolt). The relationship between ψ0 and σc for spheres is given by the Gouy-Chapman relation:26

σGC ) (ψ0/4π)(κ + 1/a)

(10)

where κ is the Debye-Hu¨ckel parameter,  is the dielectric constant of the solvent (taken as 78), and a is the dendrimer radius: 1.73 nm at acidic pH.19-20 The geometric critical surface charge density was also estimated as σg ) neR/ 4πa2, where n ) 108. Results listed in Tables 1 and 2 are also plotted as the absolute value of σGC against κ in Figure 9. Also included are results for the reverse-charged micelle-polyelectrolyte system, comprised of DMDAO (a cationic micelle with radius 2.6 nm9) and PAMPS (an anionic polyelectrolyte with charge spacing nearly the same as that of PMAPTAC). As seen in Figure 9, the dependence of σ GC on κ is linear within the rather wide limits of experimental error, and the slope is an inverse function of polyelectrolyte charge density, as expected from eq 1. In conflict with eq 1, the straight lines determined by the least-squares method do not extrapolate to σGC ) 0 at κ ) 0. However, since the number of COO- ions per dendrimer is less than 10 when σGC < 0.5, it is difficult to justify a model of a uniformly charged sphere for G3. The results obtained for six different systems, over a range of ionic strengths, are summarized in Figure 10 as the dependence of σGCξ on κ. Aside from two aberrant data points, the results for PMAPTAC and MAPTAC copolymers roughly conform to a single curve, with the results for the polyelectrolyte-micelle system forming a parallel curve, shifted toward smaller σGCξ. The large radius of DMDAO micelles relative to that of G3 may be

Figure 10. Log-log plots of σGCξ against κ in TMACl for PMAPTAC (O), MAPTAC-AAm copolymer (60%/40%) (4), MAPTAC-AAm copolymer (40%/60%) (0), MAPTAC-AAm copolymer (30%/70%) (3), MAPTAC-AAm copolymer (15%/ 85%) (?), and DMDAO-PAMPS in NaCl (b).

expected to favor polyelectrolyte binding, much as the binding of PDADMAC to G3 is stronger than its binding to G1 at equal σ;27 this leads to complex formation at lower σGCξ. The values of σGCξ for PDADMAC are much smaller, possibly reflecting the large geometric charge density of (25) Morawetz, H. Macromolecules in Solution, 2nd ed.; John Wiley & Sons: New York, 1975; pp 344-396. (26) Loeb, A. L.; Overbeek, J. Th. G.; Wiersema, P. H. The Electrical Double Layer around a Spherical Particle; MIT Press: Cambridge, U.K., 1961. (27) Zhang, H.; Ray, J.; Manning, G. S.; Moorefield, C.; Newkome, G.; Dubin, P. L. J. Phys. Chem., in press.

4250 Langmuir, Vol. 15, No. 12, 1999

PDADMAC (see above). However, most of the data obtained for G3 in the range κ ) 0.5 × 107 to 1.7 × 107 (0.025 < I